arXiv:0805.2014v2 [math.FA] 1 Sep 2008 oteeitneo eti yeo arxcle a called of type certain constru a for of 7]). existence known 17, the [11, are to e.g. means (see, few case relatively complex the 10], in 3, frames dimensi 8, r finite 18, and for the frames 13] [21, with tight comparison 12, equiangular In 4, of 7]). [1, construction [2, e.g. e.g. (see, (see, computing processing quantum signal from ranging matrix hwta h xsec fsc edlmtie sequivalen is matrices Seidel such of ma Seidel existence large the arbitrarily that find show we such particular, of In existence the roo examples. for cube conditions all sufficient are and v that necessary entries of off-diagonal number know and large eigenvalues is two arbitrarily small, an fairl is for a vertices exist and of to known number [16], are the also graphs when see graphs, 20], such has eigenvalues [19, all two matr terms have adjacency graph-theoretic matrices (Seidel) in adjacency as these interpreted when be case then can matrices ouu n.I h elcs,teeo-ignletismust entries off-diagonal these off-dia case, its real and the 0, In are entries one. diagonal modulus its self-adjoint, is it DMS-0600191. QINUA IH RMSFO OPE SEIDEL COMPLEX FROM FRAMES TIGHT EQUIANGULAR h rbe fteeitneo qinua rmsi nw t known is frames equiangular of existence the of problem The a several in role important an play frames tight Equiangular 2000 Date nti ae,w td h xsec n osrcino Sei of construction and existence the study we paper, this In h rtato a upre yNFGatDS0039 h s the DMS-0807399, Grant NSF by supported was author first The phrases. and words Key ENADG OMN,VR .PUSN N AKTOMFORDE MARK AND PAULSEN, I. VERN BODMANN, G. BERNHARD coe 3 2018. 23, October : ahmtc ujc Classification. Subject Mathematics ARCSCNANN UEROSO UNITY OF ROOTS CUBE CONTAINING MATRICES 1]wt w ievle.Amatrix A eigenvalues. two with [10] eua,drce rps ecntuteape fsc fra a such exhibit mat of vectors. also examples Seidel many construct We We complex unity. graphs. frames, vect directed of tight regular, frame roots two equiangular cube any these the between between of product multiple inner common the a existenc which the to for equivalent is frames matrices exa such have unity of of existence The roots cube containing matrices Seidel plex Abstract. edrv aiyvrfibecniin hc hrceiew characterize which conditions verifiable easily derive We qinua ih rms ot fuiy ietdgraphs. directed unity, of roots frames, tight equiangular 1. Introduction 21,5C7 05B20. 52C17, 42C15, 1 Q saSie arxpoie that provided matrix Seidel a is edlmatrix Seidel feuaglrtight equiangular of e tytoeigenvalues. two ctly e iharbitrarily with mes nlra ibr spaces Hilbert real onal cn uhrb S Grant NSF by author econd oa nre r l of all are entries gonal arcsadconstruct and matrices rcso hstp.We type. this of trices opeectlgof catalog complete y ie,adhighly and rices, so nt.W derive We unity. of ts oto eut nthe on results of host to therein) eferences oteeitneof existence the to t eso mathematics, of reas cso rps The graphs. of ices rie [5]. ertices encharacterized been r salways is ors .Mroe,such Moreover, n. tn equiangular cting relationship e arcswith matrices del 1]or [15] l be all eequivalent be o e com- hen signature ± ;such 1; 2 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE certain highly regular directed graphs, and hence, show that such directed graphs exist on an arbitrarily large number of vertices. This paper is organized as follows. We complete the introduction with a more detailed description of equiangular tight frames, Seidel matrices and their rela- tionship. In Section 2, we derive conditions for the existence of such matrices. In Section 3, we refine these conditions and use our relations, in Section 4, to deter- mine the possible sizes of such Seidel matrices for n 100. Section 5 discusses the connections with directed graphs. Section 6 contains≤ examples and Section 7 gives a method for constructing examples of arbitrarily large size.

1.1. Equiangular Tight Frames. Let be a real or complex Hilbert space. H A finite family of vectors f1,...,fn is called a frame provided that there exist strictly positive real numbers{ A and }B such that n 2 2 2 A x x,fj B x for all x . k k ≤ |h i| ≤ k k ∈H i X=1 A frame is said to be a tight frame if we can choose A = B. When A = B = 1, then the frame is called a normalized tight frame or a Parseval frame. Replacing fi by fi/√A always normalizes a tight frame. One can show (see [4, p. 21], for example) that a family f1,...,fn is a tight frame with constant A if and only if { } n 1 (1.1) x = x,fi fi for all x . A h i ∈H i X=1 There is a natural equivalence relation for tight frames, motivated by simple operations on the frame vectors which preserve identity (1.1). We say that two tight frames f1,f2,...fn and g1, g2,...gn for are uni- tarily equivalent if there exists{ a unitary operator} { U on such} thatH for all H i 1, 2,...n , gi = Ufi. We say that they are switching equivalent if there exist a∈ unitary { operator} U on , a permutation π on 1, 2,...n and a family of uni- H { } modular constants λ , λ , . . . λn such that for all i 1, 2,...n , gi = λiUf . { 1 2 } ∈ { } π(i) If is a real Hilbert space, U is understood to be orthogonal, and all λi 1 . HIn this paper we shall be concerned only with Parseval frames for the k∈-dimen- {± } sional complex Hilbert space Ck, equipped with the canonical inner product. We use the term (n, k)-frame to mean a Parseval frame of n vectors for Ck. Every such Parseval frame gives rise to an isometric embedding of Ck into Cn via the map k n V : C C , (Vx)j = x,fj for all j 1, 2,...n → h i ∈ { } which is called the analysis operator of the frame. Because V is linear, we may identify V with an n k matrix and the vectors f1,...,fn are the respective columns of V . Conversely,× given any n k matrix{V that defines} an isometry, if ∗ × we let f1,...,fn denote the columns of V ∗, then this set is an (n, k)-frame and V is the{ analysis} operator of the frame. EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 3

If V is the analysis operator of an (n, k)-frame, then since V is an isometry, we see that V V = Ik and the n n matrix VV is a self-adjoint projection of rank ∗ × ∗ k. Note that VV ∗ has entries (VV ∗)ij = ( fj,fi ). Thus, VV ∗ is the Grammian matrix (or correlation matrix) of the set ofh vectors.i Conversely, any time we have an n n self-adjoint projection P of rank k, we can always factor it as P = VV × ∗ for some n k matrix V . In this case we have V V = Ik and hence V is an × ∗ isometry and the columns of V ∗ are an (n, k)-frame. Moreover, if P = WW ∗ is another factorization of P , then there exists a unitary U such that W ∗ = UV ∗, and the frame corresponding to W differs from the frame corresponding to V by applying the same unitary to all frame vectors, which is included in our equivalence relation. However, switching equivalence is coarser than just identifying all frames with the same Grammian. In [10] it is shown that it corresponds to identifying frames for which the Grammians can be obtained from each other by conjugation with diagonal unitaries and permutation matrices.

Definition 1.1. An (n, k)-frame f ,...,fn is called uniform if there is a constant { 1 } u> 0 such that fi = u for all i. An (n, k)-frame is called equiangular if all of the frame vectors arek non-zerok and the angle between the lines generated by any pair of frame vectors is a constant, that is, provided that there is a constant b such that fi/ fi ,fj/ fj = b for all i = j. |h k k k ki| 6 Many places in the literature define equiangular to mean that the (n, k)-frame is uniform and that there is a constant c so that fi,fj = c for all i = j. However, the assumption that the frame is uniform is not|h neededi| in our definition6 as the following result shows. k Proposition 1.2. Let f ,...,fn be a tight frame for C . If all frame vectors { 1 } are non-zero and if there is a constant b so that fi/ fi ,fj/ fj = b for all |h k k k ki| i = j, then fi = fj for every i and j. 6 k k k k Proof. Without loss of generality, we may assume that the frame is a Parseval n 2 frame, so that P = ( fj,fi ) is a projection of rank k. Hence, P = P and so h i i,j=1 upon equating the (i, i)-th entry and using the fact that the trace of P is k, we see 2 n 4 n 2 2 2 4 that fi = fi,fi = fj,fi fi,fj = fi + b fi fj = fi + k k h i j=1h ih i k k j=i k k k k k k b2 f 2(k f 2), which shows that f 2 is a (non-zero)6 constant independent i i P i P ofki. k − k k k k 

In [10] a family of (n, k)-frames was introduced that was called 2-uniform frames. It was then proved that a Parseval frame is 2-uniform if and only if it is equiangular. Thus, these terminologies are interchangeable in the literature, but the equiangular terminology has become more prevalent.

1.2. Seidel Matrices and Equiangular Tight Frames. At this point we re- visit an approach that has been used to construct equiangular tight frames [15]. The previous section shows that an (n, k)-frame is determined up to unitary equiv- alence by its Grammian matrix. This reduces the problem of constructing an (n, k)-frame to constructing an n n self-adjoint projection P of rank k. × 4 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE

2 If an (n, k)-frame f ,f ,...fn is uniform, then it is known that fi = k/n { 1 2 } k k for all i 1, 2,...n . It is shown in [10, Theorem 2.5] that if f1,...,fn is an ∈ { } {k(n k) } equiangular (n, k)-frame, then for all i = j, fj,fi = cn,k = n2(n− 1) . Thus we 6 |h i| − may write q VV ∗ = (k/n)In + cn,kQ where Q is a self-adjoint n n matrix satisfying Qii = 0 for all i and Qij = 1 for all i = j. This matrix Q is× called the Seidel matrix or | associated| with6 the (n, k)-frame. The following theorem characterizes the signature matrices of equiangular (n, k)- frames. Theorem 1.3 (Theorem 3.3 of [10]). Let Q be a self-adjoint n n matrix with × Qii = 0 and Qij = 1 for all i = j. Then the following are equivalent: | | 6 (a) Q is the signature matrix of an equiangular (n, k)-frame for some k; (b) Q2 = (n 1)I + µQ for some necessarily real number µ; and (c) Q has exactly− two eigenvalues. This result reduces the problem of constructing equiangular (n, k)-frames to the problem of constructing Seidel matrices with two eigenvalues. In particular, Condition (b) is particularly useful since it gives an easy-to-check condition to verify that a matrix Q is the signature matrix of an equiangular tight frame. Furthermore, if Q is a matrix satisfying any of the three equivalent conditions in Theorem 1.3, and if λ1 < 0 < λ2 are its two eigenvalues, then the parameters n, k, µ, λ1, and λ2 satisfy the following properties:

n 1 n µn (1.2) µ = (n 2k) − = λ1 + λ2, k = − sk(n k) 2 − 2 4(n 1) + µ2 − − k(n 1) (n 1)(n k) p λ = − , λ = − − , n = 1 λ λ . 1 − n k 2 k − 1 2 r − r These equations follow from the results in [10, Proposition 3.2] and [10, Theo- rem 3.3], and by solving for λ1 and λ2 from the given equations. In the case when the entries of Q are all real, we have that the diagonal entries of Q are 0 and the off-diagonal entries of Q are 1. These matrices can be seen to be Seidel adjacency matrices of a graph [19] on±n vertices, and it has been proven that in the real case there is a one-to-one correspondence between the switching equivalence classes of real equiangular tight frames and regular two-graphs [10, Theorem 3.10]. In a similar vein, we now apply switching equivalence to complex Seidel ma- trices, in order to derive easily verifiable conditions which characterize when they have two eigenvalues.

Definition 1.4. Two Seidel matrices Q and Q′ are switching equivalent if they can be obtained from each other by conjugating with a diagonal unitary and a . EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 5

If Q is a Seidel matrix, we say that Q is in a standard form if its first row and column contains only 1’s except on the diagonal, as shown in (2.1) below. We say that it is trivial if it has a standard form which has all of its off-diagonal entries equal to 1 and nontrivial if at least one off-diagonal entry is not equal to 1. One can verify by conjugation with an appropriate diagonal unitary that the equivalence class of any Seidel matrix contains a matrix of standard form, so we only need to examine when matrices of this form have two eigenvalues. Since, in the real case, the off-diagonal entries of Q are in the set 1, 1 , it seems promising in the complex situation to harness combinatorial techniques{− } and consider the case when the off-diagonal entries of Q are mth roots of unity for some m 3. We conjecture that these cases will give a description of families of complex 2-uniform≥ frames in analogy with the characterization that has been obtained in the real case. The purpose of this paper is to examine the simplest complex case (when m = 3), and to construct new frames in this setting.

2. Signature matrices with entries in the cube roots of unity In this section we consider nontrivial signature matrices whose off-diagonal en- tries are cube roots of unity. We obtain a number of necessary and sufficient conditions for such a signature matrix of an equiangular (n, k)-frame to exist. These results are useful because they allow us to rule out many values of n and k in the search for examples of such frames. Let ω = 1 + i √3 . Then the set 1,ω,ω2 is the set of cube roots of unity. − 2 2 { } Note also that ω2 = ω and 1+ ω + ω2 = 0. Definition 2.1. We call a matrix Q a cube root Seidel matrix if it is self-adjoint, has vanishing diagonal entries, and off-diagonal entries which are all cube roots of unity. If Q has exactly two eigenvalues, then we say that it is the cube root signature matrix of an equiangular tight frame. All equivalence classes of cube root signature matrices contain representatives in standard form.

Lemma 2.2. If Q′ is an n n cube root Seidel matrix, then it is switching equiv- alent to a cube root Seidel matrix× of the form 0 1 1 1 0 ··· ··· . .∗ ···. ∗. (2.1) Q = ...... . ∗. . .  ......  . .   ∗ 1 0  ∗ ··· ∗  where the ’s are cube roots of unity. Moreover, Q′ is the signature matrix of an equiangular∗ (n, k)-frame if and only if Q is the signature matrix of an equiangular (n, k)-frame. 6 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE

Proof. Suppose that Q′ is an n n cube root Seidel matrix. Then Q′ is self-adjoint × 2 with Qij = 1 for i = j, and by Theorem 1.3 we have that (Q′) = (n 1)I + µQ′ for some| | real number6 µ. If we let U be the − 1 Q  12′  U := Q13′  .   ..     Q′   1n then U is a (since Q = 1 when i = j), and we see that Q := | ij′ | 6 U ∗Q′U is a self-adjoint n n matrix with Qii = 0 and Qij = 1 for i = j. We see that the off-diagonal elements× of Q are cube roots| of unity| and Q6 has the form shown in (2.1). (To see that the off-diagonal elements in the first row and column are 1’s, recall that Qij′ = Qji′ .) Thus Q is a cube root Seidel matrix that is unitarily equivalent to Q′. Since Q and Q′ have the same eigenvalues, if one of them is the signature matrix of an equiangular (n, k)-frame, then the same holds for the other matrix.  Next, we present the characterization of cube root Seidel matrices with exactly two eigenvalues after an elementary insight. Lemma 2.3. If a, b, c R and a1+ bω + cω2 = 0, then a = b = c. ∈ Proof. If a1+ bω + cω2 = 0, then a1+ b 1 + i √3 + c 1 i √3 = 0, and − 2 2 − 2 − 2 b c √3   b  c  hence a 2 2 + i 2 (b c) = 0. It follows that a 2 2 = 0 and b c = 0. Thus a =−b =−c. − − − −   Proposition 2.4. Let Q be a cube root Seidel matrix in standard form, and sup- pose Q satisfies the equation Q2 = (n 1)I + µQ. − n µ 2 th Then e := −3− is an integer, and for any j with 2 j n, the j column of Q th ≤ ≤ (and likewise the j row) contains e entries equal to ω, contains e entries equal 2 n+2µ+1 to ω , and contains e + µ +1= 3 entries equal to 1. Proof. For 2 j n define ≤ ≤ xj := # i : Qij = 1 { } yj := # i : Qij = ω { } 2 zj := # i : Qij = ω . { } Since the jth column of Q has n 1 nonzero entries (recall the zero on the diagonal) we have −

(2.2) xj + yj + zj = n 1. − Also, since Q2 = (n 1)I + µQ we see that for j 2 we have − ≥ 2 2 µ = µQ j = [(n 1)I + µQ] j = Q = (xj 1)1 + yjω + zjω . 1 − 1 1j − EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 7

2 Thus (xj µ 1)1 + yjω + zjω = 0, and by Lemma 2.3 we have − − (2.3) xj µ 1= yj = zj. − − Thus (2.2) becomes xj + 2(xj µ 1) = n 1 so that − − − n + 2µ + 1 (2.4) x = . j 3 and (2.3) gives n µ 2 (2.5) y = z = − − . j j 3 Since the quantities in (2.4) and (2.5) do not depend on j they are valid for any 2 column. In addition, since Q = Q∗ and ω = ω , the same equations hold for the rows of Q.  To summarize the consequences of Proposition 2.4: In the first column of Q there are n 1 entries equal to 1 and one entry equal to 0. In the jth column − n+2µ+1 n µ 2 (for j 2) there are xj = 3 entries equal to 1, there are yj = −3− entries ≥ n µ 2 2 equal to ω, there are zj = −3− entries equal to ω , and there is one entry (on the diagonal) equal to 0. Note that these values do not depend on the value of j and the same equations hold for the rows. In particular, if Q is trivial, then k = 1, and its off-diagonal entries are all 1’s, xj = n 1 and yj = zj = 0 for j 2. − ≥ 3. Equations for nontrivial cube root signature matrices Suppose that Q is a nontrivial cube root Seidel matrix. Also suppose that Q is in standard form and that Q satisfies Q2 = (n 1)I + µQ. Since Q is nontrivial − we have that Qij = ω for some 2 i, j n with i = j. For these values of i and j let ≤ ≤ 6 2 α =# k : Qik = ω and Qkj = ω { } β =# k : Qik = ω and Qkj = ω { } γ =# k : Qik = ω and Qkj = 1 { } 2 2 a =# k : Qik = ω and Qkj = ω { } 2 b =# k : Qik = ω and Qkj = ω { } 2 c =# k : Qik = ω and Qkj = 1 { } 2 A =# k : Qik = 1 and Qkj = ω { } B =# k : Qik = 1 and Qkj = ω { } C =# k : Qik = 1 and Qkj = 1 . { } We shall now establish equations relating these nine values. To begin, we see that the number of ω’s in row i is equal to α + β + γ + 1. (The +1 comes from n µ 2 the term Qij = ω.) Also, the number of ω’s in row i is equal to e := −3− by Proposition 2.4. Thus (3.1) α + β + γ = e 1. − 8 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE

In addition, the number of ω2’s in row i is equal to a+b+c, and by Proposition 2.4 the number of ω2’s in row i is equal to e. Thus (3.2) a + b + c = e. Also, the number of 1’s in row i is equal to A + B + C, and by Proposition 2.4 the number of 1’s in row i is equal to e + µ + 1. Hence (3.3) A + B + C = e + µ + 1. Next we turn our attention to the jth column. We see that the number of ω2’s in column j is equal to α + a + A, and by Proposition 2.4 the number of ω2’s in column j is equal to e. Thus (3.4) α + a + A = e. In addition, the number of ω’s in column j is equal to β + b + B + 1. (The +1 comes from the term Qij = ω.) Also, by Proposition 2.4 the number of ω’s in column j is equal to e. Thus (3.5) β + b + B = e 1. − Furthermore, the number of 1’s in column j is equal to γ + c + C, and by Propo- sition 2.4 the number of 1’s in column j is equal to e + µ + 1. Hence (3.6) γ + c + C = e + µ + 1. Finally, since Q2 = (n 1)I + µQ we have that − n 2 µω = µQij = [(n 1)I + µQ]ij = Q = QikQkj − ij Xk=1 = α(ωω2)+ β(ωω)+ γ(ω1) + a(ω2ω2)+ b(ω2ω)+ c(ω21) + A(1ω2)+ B(1ω)+ C(1)(1) = α1+ βω2 + γω + aω + b1+ cω2 + Aω2 + Bω + C1 = (α + b + C)1 + (γ + a + B)ω + (β + c + A)ω2 so that (α + b + C)1 + (γ + a + B µ)ω + (β + c + A)ω2 = 0. − It follows from Lemma 2.3 that α + b + C = γ + a + B µ = (β + c + A) and thus − (3.7) α γ a + b B + C = µ − − − − and (3.8) α β + b c A + C = 0. − − − By looking at (3.1), (3.2), (3.3), (3.4), (3.5), (3.6), (3.7), and (3.8), we have eight equations in the nine unknowns α, β, γ, a, b, c, A, B, C. These equations are not linearly independent: we see that (3.1) + (3.2) + (3.3) = (3.4) + (3.5) + (3.6). However, this is the only relation, and when we row reduce this system we obtain the seven equations 2µ 1 (Eq. 1) α B = − − 3 − 3 EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 9 2µ 4 (Eq. 2) β C = − − 3 − 3 n (Eq. 3) γ + B + C = + µ 3 µ 2 (Eq. 4) a C = − − 3 − 3 n µ 1 (Eq. 5) b + B + C = + 3 3 − 3 µ 1 (Eq. 6) c B = + − − 3 3 n 2µ 1 (Eq. 7) A + B + C = + + 3 3 3 with B and C as the two free variables. It is important to note that the above variables, α, β, γ, a, b, c, A, B, C, really should carry a subscript (i, j) since their actual values could depend on the par- ticular (i, j) that we choose satisfying Qi,j = ω and we are only asserting that for each such pair (i, j) these equations must be met, not that their values are independent of the pair that we have chosen. Similarly, we derive equations for the case that Qij = 1 for some 2 i, j n with i = j. Later, we shall prove that such an entry always exists. For th≤ese values≤ of i and6 j let 2 α′ =# k : Qik = ω and Qkj = ω { } β′ =# k : Qik = ω and Qkj = ω { } γ′ =# k : Qik = ω and Qkj = 1 { } 2 2 a′ =# k : Qik = ω and Qkj = ω { } 2 b′ =# k : Qik = ω and Qkj = ω { } 2 c′ =# k : Qik = ω and Qkj = 1 { } 2 A′ =# k : Qik = 1 and Qkj = ω { } B′ =# k : Qik = 1 and Qkj = ω { } C′ =# k : Qik = 1 and Qkj = 1 . { } We shall now establish equations relating these nine values. To begin, we see that the number of ω’s in row i is equal to α′ + β′ + γ′. Also, the number of ω’s in row i is equal to e by Proposition 2.4. Thus

(3.9) α′ + β′ + γ′ = e. 2 In addition, the number of ω ’s in row i is equal to a′ + b′ + c′, and by Propo- 2 n µ 2 sition 2.4 the number of ω ’s in row i is equal to e := −3− . Thus

(3.10) a′ + b′ + c′ = e.

Also, the number of 1’s in row i is equal to A′ + B′ + C′ + 1. (The +1 comes from the (i, j)-entry.) By Proposition 2.4 the number of 1’s in row i is equal to 10 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE e + µ + 1. Hence

(3.11) A′ + B′ + C′ = e + µ. Next we turn our attention to the jth column. We see that the number of ω2’s 2 in column j is equal to α′ + a′ + A′, and by Proposition 2.4 the number of ω ’s in column j is equal to e. Thus

(3.12) α′ + a′ + A′ = e.

In addition, the number of ω’s in column j is equal to β′ + b′ + B′. Also, by Proposition 2.4 the number of ω’s in column j is equal to e. Thus

(3.13) β′ + b′ + B′ = e.

Furthermore, the number of 1’s in column j is equal to γ′ + c′ + C′ + 1, and by Proposition 2.4 the number of 1’s in column j is equal to e + µ + 1. Hence

(3.14) γ′ + c′ + C′ = e + µ. Finally, since Q2 = (n 1)I + µQ we have that − n 2 µ = µQij = [(n 1)I + µQ]ij = Q = QikQkj − ij k=1 2 2X2 2 2 = α′(ωω )+ β′(ωω)+ γ′(ω1) + a′(ω ω )+ b′(ω ω)+ c′(ω 1) 2 + A′(1ω )+ B′(1ω)+ C′(1 1) 2 2 2 = α′1+ β′ω + γ′ω + a′ω + b′1+ c′ω + A′ω + B′ω + C′1 2 = (α′ + b′ + C′)1 + (γ′ + a′ + B′)ω + (β′ + c′ + A′)ω so that 2 (α′ + b′ + C′ µ)1 + (γ′ + a′ + B′)ω + (β′ + c′ + A′)ω = 0. − It follows from Lemma 2.3 that α′ + b′ + C′ µ = γ′ + a′ + B′ = β′ + c′ + A′ and thus −

(3.15) α′ γ′ a′ + b′ B′ + C′ =+µ − − − and

(3.16) α′ β′ + b′ c′ A′ + C′ =+µ. − − − When we row reduce this system we obtain the following equations

(Eq. 8) α′ = B′

(Eq. 9) β′ = C′ µ − (Eq. 10) γ′ = e + µ B′ C′ − − (Eq. 11) a′ = C′ µ − (Eq. 12) b′ = e + µ B′ C′ − − (Eq. 13) c′ = B′

(Eq. 14) A′ = e + µ B′ C′ − − EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 11 with B′ and C′ as the two free variables. Note that it follows from these equations that, α′ = c′ = B′, β′ = a′ = C′ µ and γ′ = A′ = b′ = e + µ B′ C′. The above equations are necessary− and sufficient to character−ize nontrivial− cube root signature matrices of equiangular (n, k)-frames.

Theorem 3.1. Let Q be a self-adjoint n n matrix with Qii = 0,Qi = Q i = 1 × 1 1 for all i and Qij a cube root of unity for all i = j, 2 i, j n with at least one 6 ≤ ≤ entry that is not equal to 1. Then Q2 = (n 1)I + µQ if and only if for each pair − i = j such that Qi,j = ω, conditions Eq. 1–Eq. 7 are satisfied and for each pair 6 i = j, 2 i, j n such that Qi,j = 1, conditions Eq. 8–Eq. 14 are satisfied, where 3e6 + µ +2=≤ n.≤ Proof. We have that Q2 = (n 1)I + µQ if and only if for every (i, j) the corre- sponding entries are equal. Equality− of the (i, i)-th entries follows from the fact that the off-diagonal entries are all of modulus one. The proof of Proposition 2.4 shows that the relationships between e, µ and n are equivalent to the condition that the (i, 1)-th entries and the (1, i)-th entries of Q2 are equal to µ which are the corresponding (i, 1)-entries and (1, i)-entries of (n 1)I + µQ. − The equations, Eq. 1–Eq. 7 are equivalent to requiring that if Qi,j = ω, then the (i, j)-th entry of Q2 is equal to µω which is the (i, j)-th entry of (n 1)I +µQ. 2 − If Qi,j = ω , then Qj,i = ω and by the last argument we have that the (j, i)- th entry of Q2 is equal to the (j, i)-th entry of (n 1)I + µQ. But since both of these matrices are self-adjoint, we have that equality− of their (j, i)-th entry implies equality of their (i, j)-th entry. Finally, if Qi,j = 1, i = j, 2 i, j n, then the equations, Eq. 8–Eq. 14, are 6 ≤ ≤ equivalent to the equality of the (i, j)-th entries of Q2 and (n 1)I + µQ.  − Remark 3.2. It is important to realize that all of our results apply only to nontrivial cube root signature matrices, because we assume that the first row contains all 1’s except for one 0, and the second row contains at least one ω. For example, 0 ω ω2 one can check that Q = ω2 0 ω is a 3 3 cube root signature matrix that ω ω2 0 ×   1 0 0 satisfies Q2 = 2I + Q. However, conjugating Q by the unitary matrix 0 ω 0 0 0 ω2 0 1 1 shows that Q has standard form 1 0 1 , and therefore Q is trivial. Thus e = 0 1 1 0 and the starting point (3.1) for deriving  our conditions does not hold.

Because α, β, γ, a, b, c, A,B,C, α′, β′, γ′, a′, b′, c′, A′,B′ and C′ are all non- negative integers, the above equations have a number of consequences for the values n and µ. We now state some results exploring these consequences. Re- markably, these consequences all seem to stem from the unprimed equations. Proposition 3.3. Let Q be a nontrivial cube root signature matrix of an equian- gular (n, k)-frame, satisfying Q2 = (n 1)I + µQ. Then the following hold: − (a) The value µ is an integer and µ 1 (mod 3). (b) The integer n satisfies n 0 (mod≡ 3). ≡ 12 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE

(c) If λ1 < 0 < λ2 are the eigenvalues of Q, then λ1 and λ2 are integers with λ1 2 (mod 3) and λ2 2 (mod 3). (d) The≡ integer 4(n 1) +≡µ2 is a perfect square and in addition we have 4(n 1) + µ2 0− (mod 9). − ≡ Proof. For (a) note that Eq. 7 shows that A + B + C = (n + 2µ + 1)/3 and hence n µ 2 A + B + C = e + µ + 1, with e = −3− . Since e is an integer by Proposition 2.4, and since A, B, and C are integers, it follows that µ is an integer. In addition Eq. 1 shows that 2µ = 3(α B) 1 so that 2µ 2 (mod 3) and µ 1 (mod 3). For (b) we see that Eq.− 3 implies− −n = 3(γ + B +≡ C µ). Since γ, B≡, and C are integers, and since µ is an integer by (a), we have that− n 0 (mod 3). ≡ n 1 Z For (c) we use (1.2) and (a) to see that (n 2k) k(n− k) = µ , and thus − − ∈ n 1 k(n 1) 2 n 1 q k(n− k) Q and λ1 = (n −k) = k k(n− k) Q. In addition, using (1.2) − ∈ − − − − ∈ weq have that λ2 = (1 n)/λq1 Q. Thus λq1 and λ2 are both rational. Because Q satisfies Q2 = (n 1)I−+ µQ, the∈ polynomial p(x)= x2 µx (n 1) annihilates − − − − Q, and hence the minimal polynomial of Q divides p(x) and the eigenvalues λ1 and λ2 are rational roots of p(x). Since the coefficients of p(x) are integers and the leading coefficient of p(x) is 1, the Rational Root Theorem tells us that the only rational roots of p(x) are integers. Hence λ1 and λ2 are integers. Finally, using (1.2) we have that λ1 + λ2 = µ, so by (a) we have (3.17) λ + λ 1 (mod 3). 1 2 ≡ Also, using (1.2) we have that λ λ = 1 n, so by (b) we have 1 2 − (3.18) λ λ 1 (mod 3). 1 2 ≡ There are only three possibilities for the residue of an integer modulo 3 (namely 0, 1, or 2) and a consideration of cases shows that the only situation in which (3.17) and (3.18) are both satisfied is when λ 2 (mod 3) and λ 2 (mod 3). 1 ≡ 2 ≡ For (d) we use (1.2) to write k = n µn . Thus 4(n 1) + µ2 = 2 − 2√4(n 1)+µ2 − µn − 2 n 2k Q by (a). Since n and µ are integers, we have thatp 4(n 1) + µ is − ∈ − an integer, and 4(n 1) + µ2 is rational if and only if 4(n 1) + µ2 is an − − integer. Thus 4(n 1) + µ2 = m for some m Z and 4(n 1) + µ2 = m2, so p − ∈ p− that 4(n 1) + µ2 is a perfect square. Furthermore, since 4(n 1) + µ2 = m2 p and we have− µ 1 (mod 3) by (a) and n 0 (mod 3) by (b),− it follows that m2 0 (mod 3).≡ Thus 3 divides m2, and since≡ 3 is prime, we have that 3 divides m.≡ Hence 9 divides m2 = 4(n 1) + µ2 and 4(n 1) + µ2 0 (mod 9).  − − ≡ Proposition 3.3 shows that µ 1 (mod 3) and n 0 (mod 3). However, we can do slightly better than this, as≡ the following propositi≡ on shows. Proposition 3.4. Let Q be a nontrivial cube root signature matrix of an equian- gular (n, k)-frame, satisfying Q2 = (n 1)I + µQ. Then one of the following three cases must hold: − n 0 (mod 9) and µ 7 (mod 9), or • ≡ ≡ EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 13

n 3 (mod 9) and µ 1 (mod 9), or • n ≡ 6 (mod 9) and µ ≡ 4 (mod 9). • ≡ ≡ Proof. From Proposition 3.3(b), we have that n 0 (mod 3) so that n is congruent to 0, 3, or 6 modulo 9. Also, from Proposition 3.3(a),≡ we have that µ 0 (mod 3) so that µ is congruent to 1, 4, or 7 modulo 9. By Proposition 3.3(d)≡ we have 4(n 1) + µ2 0 (mod 9) and by considering the possible values of n and µ modulo− 9, we see≡ the only way this equation is satisfied is if one of the three cases in the statement of this proposition holds.  Proposition 3.5. Let Q be a nontrivial cube root signature matrix of an equian- 2 n µ 2 gular (n, k)-frame, satisfying Q = (n 1)I + µQ. If we set e := −3− , then e is an integer with − e 0 (mod 3) ≡ and e satisfies 2n 4n 9 e − . 9 ≤ ≤ 9 Proof. The fact that e is an integer follows from Proposition 2.4. Also since 3e = n µ 2, by considering the three possibilities of Proposition 3.4, we see that in any− of− these three cases we have 3e 0 (mod 9). Hence e 0 (mod 3). ≡ ≡ Next we look at Eq. 1, which shows that α B = 2µ 1 . Since α is a − − 3 − 3 non-negative integer we have that 0 α = B 2µ 1 and ≤ − 3 − 3 2µ + 1 (3.19) B . ≥ 3 Also, by Eq. 2 we have that β C = 2µ 4 . Since β is a non-negative integer, − − 3 − 3 we have that 0 β = C 2µ 4 and ≤ − 3 − 3 2µ + 4 (3.20) C . ≥ 3 n µ 1 2µ+1 In addition, by Eq. 5 we have that b + B + C = 3 + 3 3 = e + 3 . Since b is a non-negative integer, we may use (3.19) and (3.20) to− obtain 2µ + 1 2µ + 1 2µ + 4 e + = b + B + C B + C + 3 ≥ ≥ 3 3 so that e 2µ+4 . Thus 3e 2µ +4=2n 6e, and ≥ 3 ≥ − 2n e . ≥ 9 For the upper bound on e, we may assume that Q is in standard form, so that every row of Q contains at least one 1, and consequently C 1. Using Eq. 3 and the fact that γ and B are non-negative integers, we have that≥ n 4n 1 γ + B + C = + µ = 3e + 2 ≤ 3 − 3 − 4n so 3e 3 3 and ≤ − 4n 9 e − . ≤ 9 14 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE



Remark 3.6. By Proposition 2.4, there are e entries that are equal to ω and e 2 entries that are equal to ω . Thus, excluding the Qi,1 entry, there are n 2 2e entries that are equal to 1. Thus, by the above inequalities, for each i, 2 − i − n, 4n 9 n ≤ ≤ there are always at least n 2 2e n 2 2 − = values of j, 2 j n for − − ≥ − − 9 9 ≤ ≤ which Qi,j = 1. Hence, there is always at least one such entry. Corollary 3.7. Let Q be a nontrivial cube root signature matrix of an equiangular 2 n µ 2 (n, k)-frame satisfying Q = (n 1)I + µQ. If we set e := −3− , then µ is an integer with − µ 1 (mod 3) ≡ and µ satisfies n n 1 µ 2. − 3 ≤ ≤ 3 − Remark 3.8. The fact that µ is an integer congruent to 1 modulo 3 is shown in n µ 2 Proposition 3.3. Using that e = −3− we may rewrite the inequality in Proposi- tion 3.5 in terms of µ to obtain 1 n µ n 2. − 3 ≤ ≤ 3 − 4. Narrowing the search for cube root signature matrices In Section 3 we derived a number of conditions that the parameters (e.g., n,k,µ,e) of a nontrivial cube root Seidel matrix must satisfy to make it the sig- nature matrix of an equiangular (n, k)-frame. These conditions allow us to rule out a number of possible values for n and k. In particular, the previous results can be incorporated in an algorithm to de- termine the possible k values for a given n.

Algorithm for deducing possible (n,k) values Begin with a value of n that is divisible by 3 (see Proposition 3.3(b)). 2n 4n 9 Step 1: Find all values of e satisfying 9 e 9− with e 1 (mod 3). (See Proposition 3.5.) ≤ ≤ ≡ Step 2: For each e from Step 1, calculate the value of µ = n 3e 2. − − Step 3: For each µ from Step 2, calculate the value of k = n µn 2 2√4(n 1)+µ2 − − (see (1.2)). The only allowable (n, k)-frames with nontrivial signature matrices are those with k equal to an integer greater than one. A necessary condition for k to be an integer is that 4(n 1) + µ2 is rational. − p Remark 4.1. One may wonder why in our algorithm we do not simply use Corol- n n lary 3.7 to first find the values of µ satisfying 1 3 µ 3 2 with µ 1 (mod 3), and then proceed directly to Step 3. This would− ≤ seem≤ to− eliminate≡ the need to calculate the value of e, and reduce our algorithm by one step. It turns out, n µ 2 however, that this is less efficient. Since e = −3− , there will in general be less EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 15 values of e found in Step 1 than there will be values of µ satisfying the condition of Corollary 3.7. To demonstrate our algorithm we go through the calculations of possible values of (n, k) for 2 k

2 1 n=3. Step 1: We need 3 e and e 3 , which cannot occur. Thus there are no allowable values of k in this≤ case which≤ lead to a nontrivial signature matrix. 1 2 n=6. Step 1: We need 1 3 e 1 3 and e 0 (mod 3), which cannot occur. Thus there are no allowable values≤ of≤ k in this case.≡ n=9. Step 1: We need 2 e 3 and e 0 (mod 3). Thus e equals 3. Step 2: For e = 3 we have µ = 2.≤ Step≤ 3: For µ ≡= 2 we have k = n µn = 6. 2 2√4(n 1)+µ2 − − − − Thus (9, 6) is an allowable (n, k) value. n=12. Step 1: We need 2 2 e 4 1 and e 0 (mod 3). Thus e equals 3. Step 2: 3 ≤ ≤ 3 ≡ For e = 3 we have µ = 1. Step 3: For µ = 1 we have 4(n 1) + µ2 = √45 / Q, which cannot occur. Thus there are no allowable values of k−in this case. ∈ 1 2 p n=15. Step 1: We need 3 3 e 5 3 and e 0 (mod 3), which cannot occur. Thus there are no allowable values≤ ≤ of k in this≡ case. n=18. Step 1: We need 4 e 7 and e 0 (mod 3). Thus e equals 6. Step 2: For ≤ ≤ ≡ e = 6 we have µ = 2. Step 3: For µ = 2 we have 4(n 1) + µ2 = √72 / Q, which cannot occur.− Thus there are no allowable− values of k−in this case. ∈ p n=21. Step 1: We need 4 2 e 8 1 and e 0 (mod 3). Thus e equals 6. Step 2: 3 ≤ ≤ 3 ≡ For e = 6 we have µ = 1. Step 3: For µ = 1 we have k = n µn = 28 / Q. 2 − 2√4(n 1)+µ2 3 ∈ Thus there are no allowable values of k in this case. − 1 2 n=24. Step 1: We need 5 3 e 9 3 and e 0 (mod 3). Thus e equals 6 or 9. Step 2: For e = 6 we have µ =≤ 4,≤ and for e = 9≡ we have µ = 5. Step 3: For µ = 4 − we have 4(n 1) + µ2 = √108 / Q, which cannot occur. For µ = 5 we have − ∈ − 4(n 1) + µ2 = √117 / Q, which cannot occur. Thus there are no allowable p values− of k in this case. ∈ p n=27. Step 1: We need 6 e 11 and e 0 (mod 3). Thus e equals 6 or 9. Step 2: For e = 6 we have µ =≤ 7, and≤ for e =≡ 9 we have µ = 2. Step 3: For µ = 7 − we have 4(n 1) + µ2 = √153 / Q, which cannot occur. For µ = 2 we have − ∈ − 4(n 1) + µ2 = √108 / Q, which cannot occur. Thus there are no allowable p values− of k in this case. ∈ p 2 1 n=30. Step 1: We need 6 3 e 12 3 and e 0 (mod 3). Thus e equals 9 or 12. Step 2: For e = 9 we have ≤µ =≤ 1, and for e ≡= 12 we have µ = 8. Step 3: For − µ = 1 we have 4(n 1) + µ2 = √117 / Q, which cannot occur. For µ = 8 − ∈ − we have 4(n 1) + µ2 = √180 / Q, which cannot occur. Thus there are no p allowable values− of k in this case. ∈ p 1 2 n=33. Step 1: We need 7 3 e 13 3 and e 0 (mod 3). Thus e equals 9 or 12. Step 2: For e = 9 we have≤ µ≤= 4, and for≡e = 12 we have µ = 5. Step 3: − 16 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE

For µ = 4 we have k = n µn = 11, which is allowed. For µ = 5 we 2 2√4(n 1)+µ2 − − − have 4(n 1) + µ2 = √153 / Q, which cannot occur. Thus (33, 11) is the only allowable (n,− k) value for n = 33.∈ p n=36. Step 1: We need 8 e 15 and e 0 (mod 3). Thus e equals 9 or 12 or 15. Step 2: For e = 9 we have≤ ≤µ = 7, for e≡= 12 we have µ = 2, and for e = 15 − we have µ = 11. Step 3: For µ = 7 we have 4(n 1) + µ2 = √189 / Q, which cannot occur.− For µ = 2 we have k = n µn− = 21, which∈ is allowed. 2 2p√4(n 1)+µ2 − − − For µ = 11 we have 4(n 1) + µ2 = √261 / Q, which cannot occur. Thus (36, 21) is− the only allowable− (n, k) value for n =∈ 36. p 2 1 n=39. Step 1: We need 8 3 e 16 3 and e 0 (mod 3). Thus e equals 9 or 12 or 15. Step 2: For e = 9 we≤ have≤µ = 10, for e≡= 12 we have µ = 1, and for e = 15 we have µ = 8. Step 3: For µ = 10 we have 4(n 1) + µ2 = √251 / Q, which − − ∈ cannot occur. For µ = 1 we have 4(n 1) + µ2 = √153 / Q, which cannot − p ∈ occur. For µ = 8 we have 4(n 1) + µ2 = √216 / Q, which cannot occur. p Thus there are no− allowable values of− k in this case. ∈ 1 p 2 n=42. Step 1: We need 9 3 e 17 3 and e 0 (mod 3). Thus e equals 12 or 15. Step 2: For e = 12 we have≤ ≤µ = 4, and for≡e = 15 we have µ = 5. Step 3: − For µ = 4 we have 4(n 1) + µ2 = √180 / Q, which cannot occur. For µ = 5 − ∈ − we have 4(n 1) + µ2 = √189 / Q, which cannot occur. Thus there are no p allowable values− of k in this case. ∈ p n=45. Step 1: We need 10 e 19 and e 0 (mod 3). Thus e equals 12 or 15 or 18. Step 2: For e = 12 we have≤ ≤µ = 7, for e≡= 15 we have µ = 2, and for e = 18 we have µ = 11. Step 3: For µ = 7 we have k = n µn − = 12, which is 2 2√4(n 1)+µ2 − − − allowed. For µ = 2 we have 4(n 1) + µ2 = √180 / Q, which cannot occur. − − ∈ For µ = 11 we have 4(n 1) + µ2 = √297 / Q, which cannot occur. Thus p (45, 12) is− the only allowable− (n, k) value for n =∈ 45. p 2 1 n=48. Step 1: We need 10 3 e 20 3 and e 0 (mod 3). Thus e equals 12 or 15 or 18. Step 2: For e = 12 we≤ have≤µ = 10, for e≡= 15 we have µ = 1, and for e = 18 we have µ = 8. Step 3: For µ = 10 we have 4(n 1) + µ2 = √288 / Q, which − − ∈ cannot occur. For µ = 1 we have 4(n 1) + µ2 = √189 / Q, which cannot − p ∈ occur. For µ = 8 we have 4(n 1) + µ2 = √252 / Q, which cannot occur. p Thus there are no− allowable values of− k in this case. ∈ p We list the possible values of (n, k) in Table 1. In addition, although we do not reproduce the calculations here, the authors have computed the possible (n, k) values from 50 to 100 as well, and these are also listed in Table 1.

5. Graphs of cube root signature matrices Arguably, the most successful means to find equiangular tight frames in the real case has been via the correspondence between graphs and signature matrices. Much of the graph-theoretic approach to real signature matrices, due largely to [15], can be repeated in our case by replacing graphs by directed graphs. In this EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 17 section we make the connection between cube root signature matrices and directed graphs explicit and describe necessary and sufficient conditions that a directed graph must satisfy in order to give rise to a cube root signature matrix in standard form. We define a one-to-one correspondence between n n selfadjoint matrices whose diagonal entries are zero and whose off-diagonal entries× are cube roots of unity with directed graphs (with no loops) on n vertices in the following manner. Given such a matrix Q, we associate its rows and columns with vertices, and an entry Qi,j = ω 2 with a directed edge from the ith to the jth vertex, while Qi,j = ω is associated with a directed edge from the jth to the ith vertex. When Qi,j =1= Qj,i, then there is no edge connecting the ith and jth vertex. We denote this directed graph by G(Q). Conversely, given any directed graph G on n vertices, and an enumeration of the vertices, we obtain such a matrix which we denote by QG and we call this matrix the Seidel of the directed graph. Given a directed graph G and vertices v and w we write v w or w v to indicate that there is a directed edge from v to w. We say that→v emits the← edge and that w receives the edge. When there is no directed edge (in either direction) between v and w, we write v w. A key element of Seidel’s≀ theory [15] was the introduction of his switching equivalence of graphs. We reformulate the equivalence relation for complex Seidel matrices for directed graphs. Suppose that we have a self-adjoint matrix Q whose off-diagonal entries are cube roots of unity and we associate a directed graph to Q as above. If we let D be the diagonal matrix satisfying dj,j = 1, j = i and di,i = ω, then the directed 6 graph G1 associated to DQD∗ is obtained from the directed graph G associated to Q by inserting a directed edge from i to j whenever i and j had no edge, • replacing a directed edge from i to j by a directed edge from j to i, and • deleting all directed edges from j to i. • We shall refer to this set of operations as the ω-switching on the ith vertex. Note that unlike Seidel’s switching for undirected graphs[15], when we switch twice on the ith vertex, we do not return to the original graph. Here when we switch three times, we return to the original directed graph. If we consider cube root signature matrices that are in standard form, then in the corresponding directed graph the first vertex neither emits nor receives any edges. Thus, without loss of generality, we may ignore this vertex and focus on the directed subgraph on m = n 1 vertices. − We now wish to describe explicitly in graph theoretical terms the directed graphs that correspond to cube root signature matrices in standard form.

Definition 5.1. We call a directed graph on m vertices e-regular, provided that each vertex emits exactly e directed edges and receives exactly e directed edges.

Given a directed graph G and vertices v and w we now wish to describe the parameters that are the natural equivalents of the parameters of Section 3. 18 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE

Given a directed graph G with vertex set V and vertices v and w with v w, we set → α =# u V : v u and w u { ∈ → → } β =# u V : v u and u w { ∈ → → } γ =# u V : v u and u w { ∈ → ≀ } a =# u V : u v and w u { ∈ → → } b =# u V : u v and u w { ∈ → → } c =# u V : u v and u w { ∈ → ≀ } A =# u V : u v and w u { ∈ ≀ → } B =# u V : u v and u w { ∈ ≀ → } C =# u V : u v and u w . 1 { ∈ ≀ ≀ } We have introduced the slight change of notation, C1, from our earlier notation C, because since we have omitted the corresponding first vertex, we will have that C1 = C 1. Similarly,− given a directed graph G with vertex set V and vertices v and w such that v w, we set ≀

α′ =# u V : v u and w u { ∈ → → } β′ =# u V : v u and u w { ∈ → → } γ′ =# u V : v u and u w { ∈ → ≀ } a′ =# u V : u v and w u { ∈ → → } b′ =# u V : u v and u w { ∈ → → } c′ =# u V : u v and u w { ∈ → ≀ } A′ =# u V : u v and w u { ∈ ≀ → } B′ =# u V : u v and u w { ∈ ≀ → } C′ =# u V : u v and u w . 1 { ∈ ≀ ≀ } Again we will have that C1′ = C′ 1. We now reinterpret Theorem 3.1− in the terminology of directed graphs. Theorem 5.2. There exists a nontrivial n n matrix Q in standard form whose off-diagonal entries are cube roots of unity× and such that Q satisfies Q2 = (n 1)I + µQ, for some µ, if and only if for some e > 0 there exists a non-empty− e-regular directed graph on m = n 1 vertices such that each pair of vertices with a directed edge from v to w satisfies,− 6e + 1 2m (Eq. d.1) α B = − − 3 6e 2m + 1 (Eq. d.2) β C = − − 1 3 EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 19 4m 9e 5 (Eq. d.3) γ + B + C = − − 1 3 3e m + 2 (Eq. d.4) a C = − − 1 3 2m 3e 4 (Eq. d.5) b + B + C = − − 1 3 3e m + 2 (Eq. d.6) c B = − − 3 (Eq. d.7) A + B + C = m 2e 1, 1 − − while each pair of vertices v and w with no edges between them satisfies

(Eq. d.8) α′ = B′

(Eq. 9) β′ = C′ + 3e + 2 m 1 − (Eq. d.10) γ′ = m 2e 2 B′ C′ − − − − 1 (Eq. d.11) a′ = C′ + 3e + 2 m 1 − (Eq. d.12) b′ = m 2e 2 B′ C′ − − − − 1 (Eq. d.13) c′ = B′

(Eq. d.14) A′ = m 2e 2 B′ C′ . − − − − 1 Moreover, in the above case we have that µ = m 3e 1 and we obtain Q from the directed graph on n 1 vertices by adjoining a− row and− column of +1’s to the adjacency matrix of the− directed graph. The results of Section 3 can now be interpreted as necessary conditions on m and e for the existence of such graphs. If G is the directed graph associated to Q and after a sequence of switchings on various vertices, we obtain a new directed graph G1, then it is easy to see that the matrix Q1 associated to G1 satisfies Q1 = DQD∗ for some diagonal matrix whose entries are cube roots of unity. In fact D can be taken to be the diagonal matrix si whose ith diagonal entry is ω where si is the number of times that we switched on the ith vertex. The proof of Lemma 2.2 shows that given a directed graph G on n vertices and a fixed vertex v, then G is switching equivalent to a unique directed graph where the vertex v is isolated, i.e., such that v neither emits nor receives any directed edges. We let Gv denote the directed graph on n 1 vertices that one obtains by deleting this isolated vertex. − Proposition 5.3. Let G be a directed graph on n vertices with adjacency matrix Q Then the following are equivalent: (1) Q2 = (n 1)I + µQ for some real number µ, − (2) for some vertex v, the graph Gv is e-regular and satisfies Eq. d.1–Eq. d.14, (3) for every vertex v, the graph Gv is e-regular and satisfies Eq. d.1–Eq.d.14. Moreover, in this case µ = n 3e 2. − − Proof. We have that Q2 = (n 1)I + µQ if and only if (D QD)2 = (n 1)I + − ∗ − µ(D∗QD) for any diagonal matrix of whose entries consist of cube roots of unity. 20 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE

These matrices represent the adjacency matrices of all directed graphs switching equivalent to G with the same enumeration of vertices. Thus, if the adjacency matrix for G satisfies the above equation, then the adjacency matrix for every directed graph switching equivalent to G satisfies the above equation. The result now follows by applying theorem 5.2.  We conclude that a necessary condition for the existence of cube root signature matrices is the existence of directed graphs without undirected edges for which all vertices have the same in-degree and out-degree e. Moreover, if there is an edge v w, then the paths between v to w of length 2 satisfy conditions (Eq. d.1-7) and→ if there is no edge between v and w, then they satisfy conditions (Eq. d.8-14). If there is only one solution for these parameters, then this implies that the graph is a directed [6].

6. Examples of cube root signature matrices with two eigenvalues 6.1. A first example. In the literature on directed strongly regular graphs, there is a graph on 8 vertices, all with same in-degree and out-degree e = 3 [14, Example 4.1]. Indeed, it turns out that the cube root Seidel matrix of this directed graph is the signature matrix of an equiangular (9, 6)-frame. This is, in fact, the unique frame up to switching, which follows from the uniqueness of directed strongly regular graphs on 8 vertices with in-degree and out-degree 3 [9]. Theorem 6.1. The matrix 011111111 1 0 1 ω ω ωω2 ω2 ω2   1 1 0 ω2 ω2 ω2 ω ω ω 1 ω2 ω 0 ω ω2 1 ω ω2   Q = 1 ω2 ω ω2 0 ω ω ω2 1    1 ω2 ω ω ω2 0 ω2 1 ω    1 ω ω2 1 ω2 ω 0 ω2 ω    1 ω ω2 ω2 ω 1 ω 0 ω2   1 ω ω2 ω 1 ω2 ω2 ω 0    is a 9 9 nontrivial cube root signature matrix of an equiangular (9, 6)-frame. Furthermore,× any 9 9 nontrivial cube root signature matrix belonging to k = 6 is switching equivalent× to Q. Proof. To see that Q is a signature matrix, one need only verify that Q2 = 8I 2Q. We prove the claimed uniqueness using a graph-theoretic argument [9]. It− will be enough to show that up to a renumbering of the vertices there is a unique directed graph on 8 vertices satisfying the conditions, Eq. d.1–Eq. d.14. To this end first note that since e = 3, there will be exactly 9 2e 2 = 1 entry in − − each row, other than Qi,1, that is equal to 1. Thus, in the directed graph on 8 vertices, each vertex will emit 3 directed edges, receive 3 directed edges, and not be connected to exactly one vertex. However, there is a unique 6-regular graph on 8 vertices. This is because each non-adjacent pair of vertices has 6 common neighbors. Thus, the graph can be EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 21 partitioned into 4 such pairs, and is seen to be the complete 4-partite graph corresponding to this partition of 8 vertices in sets of size 2. Next one finds that the equations Eq. d.1–Eq. d.14 have a unique set of solutions, α = β = a = b = c = A = 1, γ = B = C1 = 0 and α′ = γ′ = b′ = c′ = A′ = 0, β′ = a′ = 3. Thus, in the directed graph, if v w, then since β′ = 3, the three directed edges emitted by v terminate at the three≀ directed edges received by w. Similarly, since a′ = 3, the three directed edges received by v are emitted by the three vertices where the three directed edges emitted by w terminate. One now finds that there is a unique directed graph on 8 vertices satisfying these relations.  6.2. Creating new signature matrices from old. In this section we discuss ways to form new signature matrices from existing ones.

Proposition 6.2. Suppose that Q1 and Q2 are n1 n1 and n2 n2 signature matrices satisfying the equation × × 2 Q = (ni 1)Ini 2Qi, i 1, 2 i − − ∈ { } Then the matrix Q := (Q1 + In1 ) (Q2 + In2 ) In1n2 is an n1n2 n1n2 signature matrix satisfying the equation ⊗ − × 2 Q = (n n 1)In1n2 2Q. 1 2 − − Proof. We first observe that Q is self-adjoint, and since Q1 and Q2 contain unimod- ular off-diagonal entries and the I contains 1’s along the diagonal, we have that the diagonal entries of Q are zero, and the off-diagonal entries of Q are unimodular. Furthermore, 2 2 (Q + In1n2 ) = [(Q + In1 ) (Q + In2 )] 1 ⊗ 2 2 2 = (Q + In1 ) (Q + In2 ) 1 ⊗ 2 2 2 = (Q + 2Q + In1 ) (Q + 2Q + In2 ) 1 1 ⊗ 2 2 = ((n 1)In1 2Q + 2Q + In1 ) ((n 1)In2 2Q + 2Q + In2 ) 1 − − 1 1 ⊗ 2 − − 2 2 = n In1 n In2 1 ⊗ 2 = n n (In1 In2 ) 1 2 ⊗ = n1n2In1n2 . 2 2 Thus Q + 2Q + In1n2 = n1n2In1n2 , and Q = (n1n2 1)In1n2 2Q, so Q is a signature matrix. − −  Theorem 6.3. For each m N there exists a nontrivial 9m 9m cube root ∈ × signature matrix Q satisfying Q2 = (9m 1)I m 2Q. − 9 − Proof. When m = 1 it follows from Theorem 6.1 that there exists a 9 9 nontrivial 2 × cube root signature matrix Q′ satisfying (Q′) = 8I9 2Q′. To obtain matrices of size 9m 9m we iterate the construction of the preceding− proposition: We form × Q := [(Q′ + I ) . . . (Q′ + I )] I m . 9 ⊗ ⊗ 9 − 9 m times | {z } 22 BERNHARD G. BODMANN, VERN I. PAULSEN, AND MARK TOMFORDE

By considering the Kronecker product of matrices, we see that the off-diagonal entries of Q are cube roots of unity, and hence Q is a cube root signature matrix. Furthermore, we see that Q is nontrivial, because the 9 9 matrix Q′ has a nontrivial entry ω in addition to 1’s, so the Kronecker product× with itself produces at least one nontrivial entry.  Remark 6.4. One can see that Theorem 6.3 shows there is an infinite number of nontrivial cube root signature matrices of arbitrarily large size.

Table 1. Possible (n, k) values for cube root signature matrices with 2 k < n 100. ≤ ≤ n k µ e λ1 λ2 Do they exist? 9 6 -2 3 -7 5 Yes. (Theorem 6.1) 33 11 4 9 -4 8 Unknown. 36 21 -2 12 -7 5 Unknown. 45 12 7 12 -4 11 Unknown. 51 34 -5 18 -10 5 Unknown. 81 45 -2 27 -10 8 Yes. (Theorem 6.3) 96 76 -14 36 -19 5 Unknown. 99 33 7 30 -7 14 Unknown.

This table contains all possible (n, k) values for nontrivial cube root signature matrices, as determined by the algorithm in Section 4. Thus, if Q is a nontrivial cube root signature matrix of an equiangular (n, k)-frame with n 100, then n and k must be the values in one of the rows of this table. It is unknow≤ n whether there exist nontrivial signature matrices for all of the rows in the table.

References [1] A. Aldroubi, Portraits of frames, Proc. Amer. Math. Soc. 123 (1993) no. 6, 1661–1668. [2] B. G. Bodmann, D. W. Kribs, and V. I. Paulsen, Decoherence-insensitive quantum com- ∗ munication by optimal C -encoding, IEEE Transactions on Information Theory 53 (2007) no. 12, 4738–4749. [3] E. R. Berlekamp, J. H. van Lint, and J. J. Seidel, A strongly regular graph derived from the perfect ternary Golay code, A survey of combinatorial theory (J. N. Srivastava et al., ed.), North Holland, 1973, pp. 25–30. [4] P. G. Casazza, The art of frame theory, Taiwanese J. Math. 4 (2000) no. 2, 129–201. [5] A. R. Calderbank, P. J. Cameron, W. M. Kantor, and J. J. Seidel, Z4-Kerdock codes, orthogonal spreads, and extremal Euclidean line-sets, Proc. London Math. Soc. 75 (1997) no. 2, 436–480. [6] A. Duval, A directed version of strongly regular graphs, J. Combin. Theory Ser. A 47 (1988) 71–100. [7] C. Godsil and A. Roy, Equiangular lines, mutually unbiased bases and spin models, European J. Combin. In press (2008). [8] J.-M. Goethals and J. J. Seidel, Strongly regular graphs derived from combinatorial designs, Can. J. Math. 22 (1970) 597–614. EQUIANGULAR TIGHT FRAMES FROM COMPLEX SEIDEL MATRICES 23

[9] J. M. Hammersley, The friendship theorem and the love problem, Surveys in Combinatorics, Cambridge University Press, 1983, pp. 31–54. [10] R. Holmes and V. I. Paulsen, Optimal frames for erasures, Appl. 377 (2004) 31–51. [11] D. Kalra, Complex equiangular cyclic frames and erasures, Linear Algebra Appl. 419 (2006) 373–399. [12] J. Kovaˇcevi´cand A. Chebira, Life beyond bases: The advent of frames (part i), IEEE Signal Processing Magazine 24 (2007) no. 4, 86–104. [13] , Life beyond bases: The advent of frames (part ii), IEEE Signal Processing Magazine 24 (2007) no. 5, 115–125. [14] M. Klin, A. Munemasa, M. Muzychuk, and P.-H. Zieschang, Directed strongly regular graphs obtained from coherent algebras, Linear Algebra Appl. 377 (2004) 83–109. [15] P. W. H. Lemmens and J. J. Seidel, Equiangular lines, J. Algebra 24 (1973) 494–512. [16] G. E. Moorhouse, Two-graphs and skew two-graphs in finite geometries, Linear Algebra Appl. 226-228 (1995) 529–551. [17] J. M. Renes, Equiangular tight frames from Paley tournaments, Linear Algebra Appl. 426 (2007) 497–501. [18] J. J. Seidel, Strongly regular graphs with (−1, 1, 0) adjacency matrix having eigenvalue 3, Linear Algebra Appl. 1 (1968) 281–298. [19] , A survey of two-graphs, Colloquio Internazionale sulle Teorie Combinatorie (Pro- ceedings, Rome, 1973), vol. I, Accademia Nazionale dei Lincei, 1976, pp. 481–511. [20] J. J. Seidel and D. E. Taylor, Two-graphs, a second survey, Algebraic Methods in (Proc. Internat. Colloq., Szeged, 1978) (Amsterdam) (L. Lovasz and V. S´os, eds.), North-Holland, 1981, pp. 689–711. [21] J. H. van Lint and J. J. Seidel, Equilateral point sets in elliptic geometry, Indag. Math. 28 (1966) 335–348.

Department of Mathematics, University of Houston, Houston, TX 77204-3008, USA E-mail address: [email protected]

Department of Mathematics, University of Houston, Houston, TX 77204-3008, USA E-mail address: [email protected]

Department of Mathematics, University of Houston, Houston, TX 77204-3008, USA E-mail address: [email protected]